Skip to main page content
U.S. flag

An official website of the United States government

Dot gov

The .gov means it’s official.
Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

Https

The site is secure.
The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Review
. 2022 May 24;16(5):6960-7079.
doi: 10.1021/acsnano.1c09150. Epub 2022 Apr 20.

The Magnetic Genome of Two-Dimensional van der Waals Materials

Affiliations
Review

The Magnetic Genome of Two-Dimensional van der Waals Materials

Qing Hua Wang et al. ACS Nano. .

Abstract

Magnetism in two-dimensional (2D) van der Waals (vdW) materials has recently emerged as one of the most promising areas in condensed matter research, with many exciting emerging properties and significant potential for applications ranging from topological magnonics to low-power spintronics, quantum computing, and optical communications. In the brief time after their discovery, 2D magnets have blossomed into a rich area for investigation, where fundamental concepts in magnetism are challenged by the behavior of spins that can develop at the single layer limit. However, much effort is still needed in multiple fronts before 2D magnets can be routinely used for practical implementations. In this comprehensive review, prominent authors with expertise in complementary fields of 2D magnetism (i.e., synthesis, device engineering, magneto-optics, imaging, transport, mechanics, spin excitations, and theory and simulations) have joined together to provide a genome of current knowledge and a guideline for future developments in 2D magnetic materials research.

Keywords: 2D magnetic materials; CrI3; Fe3GeTe2; atomistic spin dynamics; magnetic genome; magneto-optical effect; neutron scattering; van der Waals.

PubMed Disclaimer

Conflict of interest statement

The authors declare no competing financial interest.

Figures

Figure 1
Figure 1
Timeline of developments in 2D magnets. Since early 2016, a few results on monolayer phosphides MPX3 (M = Fe, Mn, Ni, Cd; X = S, Se), and CrSiTe3 appeared in the literature, with results on electron tunneling in MnPS3 also being reported. The conclusive measurements in 2017 of magnetism on CrI3 and Cr2Ge2Te6 sparked an increasing interest in several subjects involving magnetism in 2D. Results on spin–lattice coupling collected from CrCl3 also provided different mechanisms involving vibrations and spins in 2D. In 2018, the electric control of magnetism, giant magnetoresistance, and a potential 2D magnet (i.e., VSe2) displaying room-temperature magnetism attracted substantial interest in the community. In 2019, experimental evidence of stacking-dependent magnetic properties,, pressure effects,, and giant second-harmonic generation (SHG) drove the field toward intriguing magnetic properties. In 2020, spin-textures such as skyrmions, spirals, and spin-waves indicate that topologically nontrivial spins are a reality on 2D magnets. In 2021, a few reports on twisted magnetic layers,, together with the hybrid character of narrow domain-walls on CrI3, raised possibilities for the angular control of magnetic features and domain-wall based applications (i.e., racetrack). All images adapted from the references cited above with permission as follows. Panels from (2016) reprinted with permission from ref (32), copyright 2016 American Chemical Society; ref (3), copyright 2016 Royal Society of Chemistry; ref (1), copyright 2016 American Chemical Society; and ref (4), copyright 2016 AIP Publishing and reprinted with permission under a Creative Commons Attribution (CC BY) license. Panels from (2017) reprinted with permission from ref (5), copyright 2017 Springer Nature; ref (6), copyright 2017 Springer Nature; and ref (7), copyright 2017 American Physical Society. Panels from (2018) reprinted with permission from ref (8), copyright 2018 Springer Nature; ref (9), copyright 2018 Springer Nature; ref (13), copyright 2018 AAAS; with permission under a Creative Commons CC by 4.0 license from ref (15), copyright 2018 Springer Nature; and ref (17), copyright 2018 American Chemical Society. Panels from (2019) reprinted with permission from ref (22), copyright 2019 Springer Nature; ref (23), copyright 2019 Springer Nature; ref (20), copyright 2019 AAAS; ref (21), copyright 2019 AAAS; and ref (24), copyright 2019 Springer Nature. Panels from (2020) reprinted with permission from ref (25), copyright 2020 American Chemical Society; and ref (28), copyright 2020 Springer Nature. Panels from (2021) reprinted with permission from ref (29), copyright 2021 Springer Nature; ref (30), copyright 2021 Springer Nature; and ref (31), copyright 2021 John Wiley and Sons.
Figure 2
Figure 2
Role of spin dimensionality n on the 3D-axis. (a) n = 1; 1D Ising type, where spins point in either up or down along a given direction (e.g., easy axis). (b) n = 2; 2D XY type, where spins are constrained to a given plane (easy-plane anisotropy) without any restriction on which plane (e.g., XY, XZ, YZ). (c) n = 3; 3D isotropic Heisenberg type, where spins have no constraints on the direction assuming any position along the 3D sphere.
Figure 3
Figure 3
Routes toward room-temperature FM ordering in 2D layered materials by (a) electrostatic gating, (b) tuning dimensionality, and (c) proximity effect. Representative experimental examples include: (a) ionic liquid gating in Cr2Ge2Te6, achieving a dramatic increase from 60K to 180 K in the ordering temperature. Adapted with permission from ref (86). Copyright 2020 Springer Nature. (b) Substantial Tc enhancement in Cr2Te3 flakes as the thickness is reduced, measured by anomalous Hall effect. Adapted with permission from ref (80). Copyright 2020 American Chemical Society. (c) Persistent magnetic signals up to 380 K in 4 nm-thick Fe3GeTe2 interfaced with the topological insulator Bi2Te3. Adapted with permission from ref (90). Copyright 2020 American Chemical Society.
Figure 4
Figure 4
Theoretical calculations for the Curie temperature in TMD Janus systems. (a) Diagram of the VSeTe crystal structure in top-down and cross-sectional views. (b) Temperature-dependent magnetic moment and specific heat of VSeTe, obtained via Monte Carlo simulations in a nearest-neighbor Heisenberg exchange model. Panels (a) and (b) reprinted with permission from ref (97). Copyright 2009 Royal Society of Chemistry. (c) Curie temperatures of V-, Cr- and Mn-based Janus TMDs, highlighting VSSe and VSeTe as the candidates with highest Tc. Adapted with permission from ref (96). Copyright 2018 Elsevier.
Figure 5
Figure 5
Examples of large-area 2D magnets grown by MBE. (a–d) Monolayer CrCl3 on Graphene/SiC(0001) and (e–g) few-layer Fe3GeTe2 on GaAs (111). (a) Schematic crystal structure of CrCl3/graphene/6H-SiC layers in top-down view and cross-section view. (b) In situ RHEED pattern of the substrate and monolayer CrCl3 grown by MBE, along Γ–M of graphene (Γ–K of SiC). Streaks from different high-symmetry directions of CrCl3 are observed, implying a twisted in-plane orientation of the grains. (c) STM topography of a monolayer CrCl3 grown on graphene/6H-SiC(0001), indicating a homogeneous coverage on long length scales. Inset: A magnified topography image, which reveals the grain boundaries. (d) Atom resolved image of the CrCl3 lattice featuring a moiré pattern (upper panel) and its Fourier transformed image (lower panel). Panels (a–d) adapted with permission from ref (78). Copyright 2021 AAAS. (e) Crystal structure of Fe3GeTe2. (f) RHEED oscillations indicating layer-by-layer growth of Fe3GeTe2 (0001) on GaAs (111), and the corresponding electron diffraction pattern (inset). The inferred growth rate is 111 s per monolayer. (g) Transmission electron microscopy of a Fe3GeTe2/GaAs cross-section, indicating the (111)/(0002) epitaxial relationship. Panels (e–g) adapted with permission under a Creative Commins CC BY license from ref (99). Copyright 2017 Springer Nature.
Figure 6
Figure 6
Enormous TMR in vdW MTJs induced by layered antiferromagnetism. (A) Reflective magnetic circular dichroism (RMCD) of bilayer (left) and trilayer (right) CrI3 showing layer-dependent magnetism. Reproduced with permission from ref (13). Copyright 2018 AAAS. (B) Magnetic-field-dependent current change in a vdW MTJ incorporating magnetic CrI3. (C) Schematic energy diagrams of CrI3-based MTJs with AF barrier (top) and FM barrier (bottom). Panels (B and C) are adapted with permission from ref (16). Copyright 2018 American Chemical Society. (D) Summary of TMR values as a function of thickness. Reproduced from ref (112). Copyright 2019 American Chemical Society.
Figure 7
Figure 7
Summary of intrinsic magnetoresistance in vdW magnets. (A) Magnetoresistance ratio MRR(B) formula image in bulk CrSBr versus magnetic field (parallel to the b-axis) at various temperatures. Each MRR curve is offset for clarity. The solid black line is an MRR curve taken near the Néel temperature. The antiferromagnetic (AF), fully polarized (FP), and paramagnetic (PM) phases are labeled, and the phase boundary is denoted by dashed black lines. Schematics showing the orientation of the spins in the AF and FP states are given above the plot. Reproduced with permission from ref (110). Copyright 2020 John Wiley and Sons. (B) Ryx of a 5-layer MnBi2Te4 sample as a function of external magnetic field applied perpendicular to the sample plane at T = 1.6 K. Data are symmetrized to remove the Rxx component. (C) Rxx of a 5-layer MnBi2Te4 flake as a function of magnetic field acquired at various temperatures. Data are symmetrized to remove the Ryx component. Inset shows the layered crystal structure of MnBi2Te4 in the AF state. Panels (B) and (C) are reproduced with permission from ref (111). Copyright 2020 AAAS. (D) Ball and stick model of the Cr2Ge2Te6 crystal structure. (E) Magnetoresistance formula image curves for T = 60 K and back-gate voltage of 3.9 V for a 22 nm-thick Cr2Ge2Te6 flake. The background is removed for clarity. The magnetic field is applied in the out-of-plane direction. Unprocessed data are shown in the inset. Panels (D) and (E) are reproduced with permission from ref (86). Copyright 2020 Springer Nature. (F) Side view of the atomic lattice of bilayer Fe3GeTe2. The dashed rectangular box denotes the crystal unit cell. (G) Temperature-dependent magnetic field (out-of-plane) sweeps of the Hall resistance measured on a 12 nm thick Fe3GeTe2 flake. Panels (F) and (G) are reproduced with permission from ref (77). Copyright 2018 Springer Nature.
Figure 8
Figure 8
Comparison between tunnelling differential conductance spectra between Gr/CrI3/Gr and Gr/CrBr3/Gr devices in a magnetic field applied out-of-plane of the magnetic film for CrI3 sample and in-the-plane of the magnetic film for CrBr3 sample. For both materials, narrow low energy resonances appear, which are dispersive in the magnetic field. The color maps demonstrating the dependence of the differential conductance for (a, b) Gr/CrI3/Gr and (c) Gr/CrBr3/Gr devices illustrate the evolution of such resonances. The quantitative parameters describing the magnonic states, i.e., their zero-field energy and the slope ΔEB can be extracted based on the linear fits to the evolution of the energy of the resonances in the magnetic field (b, d). Panels (a) and (b) are adapted with permission from ref (14). Copyright 2018 AAAS. Panels (c) and (d) are adapted with permission from ref (131). Copyright 2018 Springer Nature.
Figure 9
Figure 9
Temperature evolution of the differential conductance spectra in an out-of-plane magnetic field of 17.5 T is indicative of elastic and inelastic tunnelling processes mediated by magnons. The experimental data (a) shows qualitative agreement with theoretical predictions (b), which consider inelastic scattering processes with an emission of a magnon and elastic tunnelling processes involving two magnons. Both processes can be distinguished by applying a bias which shifts the mutual alignment of Landau levels (LLs) in the graphene electrodes (c, d). All panels adapted with permission from ref (131). Copyright 2018 Springer Nature.
Figure 10
Figure 10
(a) Schematic illustration in a single-particle picture of the direct band gap edge states for a W-based TMD (bottom) in the degenerate but inequivalent corners of the hexagonal Brillouin zone (top). The red dashed (blue solid) lines depict spin-up (down) band-edge states. Up (down) short arrows indicate spin-up (down) conduction band and valence band electrons. Long vertical arrows represent spin-allowed optical transitions with σ+ (red) and σ (blue) polarization. Adapted with permission from ref (146). Copyright 2018 American Physical Society. (b) Schematic illustration of the optical response of an ideal 2D semiconductor, showing the exciton ground (n = 1) and excited state resonances (n = 2, 3, 4, ...) below the renormalized quasiparticle band gap. Adapted with permission under a Creative Commons CC BY 4.0 license from ref (147). Copyright 2017 Springer Nature. The top right inset shows the energy level scheme of the exciton states, designated by their principal quantum number n, with the binding energy of the exciton ground state below the free-particle (FP) band gap. Adapted with permission from ref (146). Copyright 2018 American Physical Society. (c) Schematic energy level diagram depicting the three contributions to the valley Zeeman shifts of the band-edge states: spin (black), valley (green) and atomic orbital (purple). The dashed (solid) lines represent the energies of the states before (after) applying a positive magnetic field perpendicular to the material interface. Adapted with permission from ref (148). Copyright 2015 Springer Nature. (d) Schematics of a typical microscope for optical spectroscopy of 2D materials in epifluorescence geometry. The 2D materials can be studied at temperatures T = 4–300 K by placing them on nonmagnetic nanopositioners inside a cryostat. A solenoid allows the application of magnetic fields (B) perpendicular to the crystal plane (Faraday geometry). The excitation and collection paths can feature several polarizing optical components for PL and reflectance experiments in circular (σ) and linear (π) bases. Adapted with permission from ref (149). Copyright 2017 Springer Nature. (e) PL spectrum of ML MoSe2 at T = 4 K under continuous-wave laser excitation with 2.33 eV. The spectrum shows the neutral exciton (X0) and the lower-energy negatively charged exciton (X). The X shows a binding energy of about 30 meV (see inset). Adapted with permission from ref (150). Copyright 2013 Springer Nature. (f) Derivative of the reflectance contrast spectrum (d/dE) (ΔR/R) for a WS2 monolayer (ML). The exciton states are labeled by their respective quantum numbers. The inset shows the as-measured reflectance contrast ΔR/R for comparison. Adapted with permission from ref (151). Copyright 2014 American Physical Society.
Figure 11
Figure 11
(a) Schematic band structure depicting intravalley (left) and intervalley (right) interactions between an exciton and a spin-polarized 2DES in a lightly doped W-based ML TMD. Adapted with permission from ref (170). Copyright 2020 AIP Publishing. (b) Voltage-gate-dependent color-scale map presenting reflectance contrast spectra measured in an hBN-encapsulated ML MoSe2. Attractive and repulsive exciton polarons are visible in both n- and p-doped regimes. Adapted with permission under a Creative Commons CC BY 4.0 license from ref (171). Copyright 2019 American Physical Society. (c) Gate-voltage dependencies of the line widths (top panel) and energies (middle panel) of the exciton (red) and attractive polaron (blue) resonances on the hole-doped side of an hBN-encapusulated ML MoSe2 at B = 16 T. Bottom panel: Gate voltages corresponding to integer filling factors determined based on the positions of the local minima of the exciton line width. Adapted with permission under a Creative Commons CC BY 4.0 license from ref (171). Copyright 2019 American Physical Society. (d) Gate-voltage-dependence derivative of the reflectance contrast spectrum with respect to the gate voltage of a charge-tunable, dual-graphene-gated, and fully hBN-encapsulated MoSe2 ML (right panel). The weak, higher-energy resonance is due to Umklapp scattering of the excitons off the electron Wigner crystal. Left panel: Schematics of the exciton dispersion in a ML TMD semiconductor hosting an electron system in various structural phases. The parabolic- and linear-in-momentum exciton branches arise from the splitting of the exciton branches due to the electron–hole exchange interaction, and correspond to the exciton dipole oriented along transverse (T) or longitudinal (L) directions with respect to the momentum vector, respectively. Adapted with permission from ref (172). Copyright 2021 Springer Nature. (e) Top panels: schematics of a Wigner crystal in a MoSe2 bilayer with an intercalated 1 nm thick layer of hBN (left). The top right panels show schematics of commensurate stacking in bilayer Wigner crystals with triangular lattices for filling ratios nt:nb of 1:1, 4:1, and 7:1, with nt and nb the carrier density in the top (blue dots) and bottom (red dots) MoSe2 layers. Bottom panels: 2D map of δ(nt, nb) as a function of total carrier density n and temperature T for nt:nb = 1:1 (right). The Wigner crystal forms in the region δ > 0 region. Theoretical schematic phase diagram of a bilayer Wigner crystal, showing both quantum and thermal phase transitions (left). Adapted with permission from ref (173). Copyright 2021 Springer Nature.
Figure 12
Figure 12
(a) Schematic illustration of a 2D triangular moiré superlattice resulting from stacking two TMD MLs with different lattice constants and/or twist angle. The moiré potentials (empty blue circles) can be loaded with either electrons or holes with spin up (red arows) or down (blue arrows). The on-site Coulomb interaction energy U and hopping amplitude t between spins in the lattice is highly tunable by the stacking angle and choice of 2D materials, enabling the investigation of the Fermi–Hubbard model. (b) Sketch of the type-II band structure of a WSe2/WS2 heterobilayer, where K and K′ represent two valley degrees of freedom. Up (down) arrows denote the spin up (down) direction. (c) Dependence of the magnetic susceptibility χ ∝ gg0 (left axis, black filled symbols) and Weiss constant θ (right axis, red empty symbols) on the filling factor of WSe2/WS2 at 1.65 K. Panels (b) and (c) adapted with permission from ref (197). Copyright 2020 Springer Nature. (d) Optically detected resistance and capacitance signal at 1 kHz (gray) and 30 kHz (black) from charge-neutral to moderate hole doping in WSe2/WS2, showing gap-like features at hole doping levels of n = n0/3 (orange dashed line), n = 2n0/3 (green dashed line) and n = n0 (blue dashed line). Adapted with permission from ref (198). Copyright 2020 Springer Nature. (e) Abundance of insulating states in WSe2/WS2 as revealed by the blue-shifts of the 2s exciton resonance in the reflectance contrast of a ML WSe2 sensor placed in close proximity to the heterobilayer (top panel). The top axis shows the proposed filling factor for the insulating states, with the corresponding configurations schematically shown in the bottom panels. Adapted with permission from ref (199). Copyright 2020 Springer Nature.
Figure 13
Figure 13
(a) Top: Schematic illustration of electron–hole pairs forming 1s and 2s excitonic states in a 2D dielectric slab. Adapted with permission from ref (151). Copyright 2014 American Physical Society. Bottom: Theoretically calculated energies of the bandgap and exciton states in a WS2 ML as a function of inverse squared external dielectric constant. Shaded areas indicate fluctuations from variations of the external screening. Adapted with permission from ref (202). Copyright 2019 Springer Nature. (b) Schematic of a device heterostructure to demonstrate the sensing capabilities of excitons in ML WSe2. (c) Reflectance contrast of the 1s (left) and 2s and 3s (middle) excitonic transitions in the WSe2 sensor, and the 1s exciton in the WS2 sample as a function of the applied gate voltage (i.e., the electron concentration in WS2). (b) and (c) Adapted with permission from ref (199). Copyright 2020 Springer Nature. (d) Left: Reflective magneto-circular dichroism as a function of magnetic field (bottom) measured in a monolayer WSe2 and trilayer CrI3 heterostructure, depicted above schematically. Orange and green curves represent magnetic field sweeping up (increase) and down (decrease), respectively. Right: Schematic of a ML WSe2 and bilayer CrI3 heterostructure (top). The layered AF spatial domains that are indistinguishable by reflective magneto-circular dichroism can be resolved by circular polarization-resolved PL from WSe2 (bottom). Adapted with permission from ref (203). Copyright 2020 Springer Nature. (e) Measurement of the CrBr3 magnetization hysteresis using the MOKE (bottom) in a device like schematically shown on top. Adapted with permission under a Creative Commons CC BY 4.0 license from ref (204). Copyright 2020 American Physical Society.
Figure 14
Figure 14
(a) Top view of CrI3 monolayer with gray and purple spheres representing Cr and I atoms. Adapted with permission from ref (5). Copyright 2017 Springer Nature. (b, c) Lateral view of bulk CrI3 in the monoclinic (b) and rhombohedral (c) phase. (d, e) Polarization resolved Raman spectra for bulk CrI3 taken at 5 K (d) and 280 K (e). (f, g) Polarization resolved Raman spectra for thin layer CrI3 taken at 5 K (f) and 280 K (g). Panels (b–g) adapted with permission under a Creative Commons CC BY license from ref (224). Copyright 2019 IOP Publishing.
Figure 15
Figure 15
(a) Raman spectra taken on bulk CrI3 in LL and LR channels at 10 K and 0 T. P(Ag), P(Eg), and M1,2 label phonon modes of Ag, Eg symmetry and coupled with layered magnetism, respectively. (b) Color map of magnetic field dependent Raman spectra taken in the LL channel. (c) Raman spectra taken on bulk CrI3 in LL and LR channels at 7 T. (d–f) Magnetic field dependence of selected Ag and Eg phonon modes. (g) A schematic illustration of the monoclinic distortion across the layered surface AF to FM transition. Adapted with permission under a Creative Commons CC BY 4.0 license from ref (218). Copyright 2020 American Physical Society.
Figure 16
Figure 16
(a) Raman spectra taken in both linear parallel and crossed channels in monolayer CrI3 at 15 K and 0 T. Polar plots of polarization dependent Ag mode intensity (∼129 cm–1) at 60 and 15 K, above and below the magnetic onset temperature TC = 45 K, respectively. (b) Raman spectra of 2L CrI3 at 15 K and at 0 and 1.5 T in linearly parallel and crossed channels. Panels (a, b) adapted with permission from ref (221). Copyright 2020 Springer Nature. (c) Raman spectra of 2L CrI3 at 1.7 K and at 0 T and −1 T in linearly parallel and crossed channel. Adapted with permissions from ref (222). copyright 2020 American Chemical Society. (d) Raman spectra of 10L CrI3 at 9 K and at selected magnetic fields. Adapted with permission under a Creative Commons CC BY license from ref (223). Copyright 2020 Springer Nature.
Figure 17
Figure 17
(a) Raman spectra taken on 1–4L CrI3 in both linear parallel and crossed channels at 10 K and 0 T. Adapted with permission under a Creative Commons CC-BY license from ref (219). Copyright 2017 National Academy of Sciences. (b) A summary of linear chain model calculation results and selection rule analysis for 1–4L CrI3.
Figure 18
Figure 18
(a) Raman spectra taken on 4L CrI3 in the RR channel at 10 K and at 0, 1, and 2 T, corresponding to the three magnetic field ranges B < Bc1, Bc1 < B < Bc2, and B > Bc2, respectively. (b) Magnetic field dependence of three modes U1,3,4. (c) A summary of the analysis of the layered magnetism-assisted Raman scattering for four phonon modes and three layered magnetic structures. (d) List of phonon modes in 4L CrI3 that contribute to the layered magnetism-assisted scattering channel with antisymmetric Raman tensors. (e) Calculated magnetic field dependence of the four phonons of 4L CrI3, U1–4, in the RR channel. Adapted with permission under a Creative Commons CC-BY license from ref (219). Copyright 2017 National Academy of Sciences.
Figure 19
Figure 19
(a) Raman spectra of 2L CrI3 taken at 10 and 70 K with an excitation wavelength of 633 nm. Solid orange lines are fits to the raw Raman spectra, using a sum of N Lorentzian profiles and a constant background, formula image. (b) Histogram plot of the fitted Lorentzian mode intensity (AN) as a function of N at 10 and 70 K. Solid curves are fits of the peak intensity profiles to the Poisson distribution functions, formula image. (c) Plot of 2D e-ph coupling constant (α2D) as a function of temperature. The dashed vertical line marks the magnetic onset TC = 45 K. Adapted with permission under a Creative Commons CC BY license from ref (220). Copyright 2020 Springer Nature.
Figure 20
Figure 20
(a) Raman spectra of 2L CrI3 acquired at 10 K in the RR and LL polarization channels with an applied out-of-plane magnetic field (B) of 1 T (top), 0 T (middle), and −1 T (bottom), respectively. (b, c) Plots of the Poisson fit amplitude A0 (b) and the 2D e-ph coupling strength α2D (c) as a function of the applied B in the RR (orange data points) and LL (blue) channels. Solid lines are step (orange and blue in b) and linear (gray in c) function fits to the magnetic field dependence of A0 and α2D, respectively. Adapted with permission under a Creative Commons CC BY license from ref (220). Copyright 2020 Springer Nature.
Figure 21
Figure 21
Birefringence effects in atomically thin CrI3. (a) The measurement of Kerr rotation angle in external magnetic field reveals finite magnetization in CrI3 monolayers. The inspection of CrI3 multilayers demonstrates AF interlayer coupling in (b) bilayers and (c) trilayers. Panels (a–c) adapted with permission from ref (5). Copyright 2017 by Springer Nature. (d, e) Qualitatively, the same behavior for atomically thin layers is observed when inspecting MCD. Adapted with permission from ref (13). Copyright 2018 AAAS.
Figure 22
Figure 22
Photoluminescence bands in CrI3 monolayers below Curie temperature display large degree of circular polarization which can be linked to the emergence of magnetization. A rapid reorientation of magnetic moments leads to the change in the sign of the polarization degree combined with a hysteric behavior with a coercive field of about 0.1 T, in agreement with analogous observation by birefringence effects. Adapted with permission from ref (215). Copyright 2017 Springer Nature.
Figure 23
Figure 23
(a) MOKE maps of a CrI3 monolayer at external magnetic fields of 0, 0.15, and 0.3 T. Adapted with permission from ref (5). Copyright 2017 Springer Nature. (b) Kerr rotation signal for Cr2Ge2Te6 bilayer flake under 0.075 T as the temperature decreases from 40 to 4.7 K. The average background signal has been subtracted and the signals are truncated at 30 μrad. Adapted with permission from ref (6). Copyright 2017 Springer Nature. (c) Pump-induced Kerr rotation as a function of pump–probe delay time in bilayer CrI3 under different in-plane magnetic fields. Adapted with permission from ref (28). Copyright 2020 Springer Nature. (d) (iii) Illustration of the wide-field MCD experimental setup. Blue and red beams represent illumination light from the laser and scattered light from the sample with different effective numeric apertures. HWP: Half-wave plate. QWP: Quarter-wave plate. (ii–iv) Optical microscopy image (ii) and polarization-enhanced MCD image (iii, iv) of a monolayer CrBr3 (white dashed box). The MCD image shows giant optical contrast of ±60% for the positive (iii) and negative (iv) remnant magnetization. Adapted with permission from ref (246). Copyright 2020 Springer Nature.
Figure 24
Figure 24
(a) STXM images show the initial labyrinth domain states stabilized at 120 and 100 K transformed to the magnetic skyrmion lattices induced by bipolar pulse bursts. The other two images at Bz = 0 mT were acquired after removing the external fields. (b) Lorentz TEM images of magnetic skyrmion lattices taken at the sample tilting angle of −20° (left), 0° (middle) and 20° (right) with respect to x-axis as illustrated in the upper panels at Bz = −40 mT and 160 K, respectively. Panels (a) and (b) are adapted with permission from ref (26). Copyright 2021 American Physical Society. (c) (i) Overfocused Lorentz-TEM images of the skyrmion bubbles taken at 93 K and in zero-field. (ii) An enlarged in-plane magnetization distribution map obtained by transport of intensity equation analysis for a selected skyrmion bubble indicated by the white dotted box in (i). (iii) Simulation of skyrmion lattices at an out-of-plane field of 60 mT. Adapted with permission from ref (25). Copyright 2020 American Chemical Society. (d) Schematic diagram of a Néel-type skyrmion on a tilt sample for Lorentz TEM imaging (left). The orange and blue circles are for positive and negative magnetizations along z direction, respectively. Brown arrows indicate the in-plane magnetization component, while gray arrows indicate the Lorentz force. Lorentz TEM images (right) observation of skyrmion lattice from under focus to over focus on WTe2/40L Fe3GeTe2 samples at 180 K with a field of 51 mT. Adapted with permission under a Creative Commons CC BY license from ref (27). Copyright 2020 Springer Nature.
Figure 25
Figure 25
(a) STM images of a CrBr3 film with adjacent monolayer (1L) and bilayer (2L) regions and spin-polarized tunneling signals on the bilayer regions as a function of magnetic field. Adapted with permission from ref (20). Copyright 2019 AAAS. (b) Antiferromagnetic domain patterns in MnBi2Te4 measured with MFM. Adapted with permission from ref (253). Copyright 2020 American Chemical Society. (c) Schematic illustration of the pulled quartz tube with two Nb or Pb superconducting leads connected to Au electrodes and image of a Nb SQUID-on-tip device. Inset: Magnified view showing the superconducting loop on the apex of the tip. The bridges that reside in the gap regions between the leads form the two weak links of the SQUID. Adapted with permission from ref (254). Copyright 2013 Nature Springer. (d) Gradient magnetometry signal associated with the fully polarized twisted bilayer graphene device. Adapted with permission from ref (255). Copyright 2021 AAAS. (e–g) Magnetic images obtained with scanning NV magnetometry. (e) Magnetization image of bilayer/trilayer CrI3 adapted with permissions from ref (21). Copyright 2019 AAAS. (f) The magnetic domains in bilayer CrBr3 adapted with permission under a Creative Commons CC BY license from ref (256). Copyright 2021 Springer Nature. (g) Stray field map of a magnetic skyrmion in the CoFeB system adapted with permission under a Creative Commons CC BY license from ref (257). Copyright 2018 Springer Nature.
Figure 26
Figure 26
(a) Crystal structure of Cr2Ge2Te6. (b) Temperature dependence of magnetization for Cr2Ge2Te6 measured in H = 1 kOe. Inset: Field dependence of magnetization at T = 2 K. (c) 2D Ising model plot of isotherms around Tc. (d) Modified Arrott plot of M1/βversus (H/M)1/γ with β = 0.194 and γ = 1.36. The straight line is the linear fit of isotherm at T = 62.5 K. (e) Temperature dependence of the spontaneous magnetization Ms (left) and the inverse initial susceptibility formula image (right) with solid fitting curves. (f) Kouvel–Fisher plots of formula image (left) and formula image (right) with solid fitting curves. (g) Isotherm MversusH plot collected at Tc = 62.7 K. Inset: The same plot in log–log scale with a solid fitting curve. (h) Scaling plots of renormalized magnetization mversus renormalized field h below and above Tc for Cr2Ge2Te6. Inset: The rescaling of the M(H) curves by MH–1/δversus εH–1/(βδ). All panels adapted with permission from ref (279). Copyright 2017 American Physical Society.
Figure 27
Figure 27
Temperature dependence of (a) |ΔSM| and (b) ΔCp in different fields with Hc; Field dependence of parameters from |ΔSM(T)| with the fitted curves: (c) formula image, (d) δTfwhm, and (e) RCPversusH with the fitted curves; (f) Modified Arrott plot based on the obtained critical exponents. Scaling of the |ΔSM(T, H)| curves: (g) normalized ΔSM(T, H) as a function of θ (inset gives Tr1 and Tr2 as a function of H); (h) – ΔSM/H(1−α)/Δversus ε/H1/Δ. All panels adapted with permission from ref (280). Copyright 1998 American Physical Society.
Figure 28
Figure 28
(a) Schematics illustrating the combination of two structural motifs, i.e., that of transition-metal dichalcogenides TMC2 (TM, transition metal; C, chalcogen) together with that of body-centered cubic iron to form Fe-rich vdW coupled ferromagnets of composition FexGeC2. (b) Three stable structures within the FenGeTe2 series with the values n = 3, 4, and 5 which were previously identified through ab initio calculations. Fe–Fe dumbbells are the common structural units that form multiple-layer Fe-rich slabs stacked through vdW-like coupling. As n increases, the number of nearest Fe neighbors per Fe atom gradually increases, which is expected to enhance the pair exchange interaction and, thus, Tc. All panels are adapted with permission under a Creative Commons CC BY-NC 4.0 license from ref (88). Copyright 2020 AAAS.
Figure 29
Figure 29
(a) Magnetic susceptibility χ as a function of the temperature T for a Fe3GeTe2 single-crystal measured under zero field-cooled (black markers) and field cooled (red markers) conditions. Open markers correspond to the inverse of χ where the blue line is a linear fit. The Curie temperature exceeds 200 K. Adapted with permission from ref (327). Copyright 2017 American Physical Society. (b) χ as a function of T for a Fe4GeTe2 single-crystal and for fields along the c-axis (solid green makers) and the ab-plane (open green markers). Solid black markers depict both the resistivity and its derivative indicating a Tc of ∼270 K with another spin-reorientation transition above 100 K. Adapted with permission under a Creative Commons CC BY-NC 4.0 license from ref (88). Copyright 2020 AAAS. (c) Magnetization as a function of the temperature for Fe5GeTe2 for fields along the c-axis (open markers) and the ab-plane (solid markers) and also polycrystalline material (gray line). It displays a Tc in excess of 280 K. Adapted with permission from ref (87). Copyright 2019 American Chemical Society.
Figure 30
Figure 30
(a) Atomic structure of monolayer Fe3GeTe2. The left panel shows the view along [001]; the right panel shows the view along [010]. Bulk Fe3GeTe2 is a layered crystal with an interlayer vdW gap of 2.95 Å. FeI and FeII represent the two inequivalent Fe sites in the +3 and +2 oxidation states, respectively. (b) Optical image of typical few-layer flakes exfoliated on an Al2O3 thin film. (c) Atomic force microscopy image of the area marked by the square in (b). Mono- and few-layer flakes are clearly visible. Scale bar, 2 μm. (d) Cross-sectional profile of the Fe3GeTe2 flakes along the white line in (c). The steps are 0.8 nm in height, or consistent with the thickness (0.8 nm) of monolayer (1L) Fe3GeTe2. (e) Normalized remanent anomalous Hall resistance formula image as a function of temperature obtained from Fe3GeTe2 thin-flake samples with varying numbers of layers. Arrows mark the FM transition temperature Tc. (f) Phase diagram of Fe3GeTe2 as layer number and temperature are varied. Tc values are determined from anomalous Hall effect, Arrott plots and RMCD are displayed in blue, red and magenta, respectively. (g) Remanent RMCD signal as a function of temperature for a sequence of selected few-layer flakes (1 L, monolayer; 2 L, bilayer; 3 L, trilayer; 4 L, four layers; 5 L, five layer). The solid lines are least-squares criticality fits of the form α(1 – T/Tc)β. Inset: derived values of the exponent β plotted as a function of thickness. (h) Thickness-temperature phase diagram. PM denotes the region in which the flake is paramagnetic, FM1 that in which it is FM with a single domain and FM2 that in which the flake exhibits labyrinthine or stripe domains. The transition temperatures, Tc, Tc1, and Tc2, are based on the temperature-dependent RMCD or anomalous Hall effect measurements for each flake thickness. The red dashed line denotes the critical thickness at which a dimensional crossover occurs. All panels are adapted with permission from ref (12). Copyright 2018 Springer Nature.
Figure 31
Figure 31
Principle of a μSR experiment. (a) Overview of the experimental setup. Spin polarized muons with spin Sμ antiparallel to the momentum pμ are implanted in the sample placed between the forward (F) and the backward (B) positron detectors. A clock is started at the time the muon goes through the muon detector (M) and is stopped as soon as the decay positron is detected in the detectors F or B. Adapted with permission from ref (348). Copyright Swiss Physical Society. (b) The number of detected positrons NF and NB as a function of time for the forward and backward detector, respectively. Reproduced with permissions from ref (349). Copyright University of Zurich. (c) The so-called asymmetry (or μSR) signal is obtained by essentially building the difference between NF and NB (eq 2). All panels are adapted with permission under a Creative Common CC BY license from ref (350). Copyright 2019 MDPI.
Figure 32
Figure 32
(a-c) Schematic illustration of the magnetically homogeneous (i.e., full volume magnetic) (d), inhomogeneous (full volume magnetic, but with domains) and phase separated (i.e., part of the volume magnetic and part paramagnetic) polycrystalline samples and the corresponding μSR spectra. The 1/3 nonoscillating μSR signal fraction originates from the spatial averaging in powder samples where 1/3 of the magnetic field components are parallel to the muon spin and do not cause muon spin precession. (d, e) Isotropic Gaussian field distribution for polycrystalline sample. Panels (a–c) adapted with permission under a Creative Common CC BY license from ref (350). Copyright 2019 MDPI.
Figure 33
Figure 33
Types of intrinsic disorder in TMDs: (a) vacancy, (b) antisite, (c) substitution. Panels (a–c) adapted with permission from ref (380). Copyright 2019 Springer Nature. (d) Schematic diagram of the phase incorporation strategy to achieve ferromagnetism of 2H-MoS2 nanosheets. Adapted with permission from ref (381). Copyright 2015 American Chemical Society.
Figure 34
Figure 34
(a) ZF μSR time spectra for the single crystal samples of Td-MoTe2 and 2H-MoTe2 recorded at T = 5 K. Adapted with permission under a Creative Common CC BY license from ref (350). Copyright 2019 MDPI. (b) Temperature dependence of the internal field Hint of 2H-MoTe2, 2H-MoSe2 as a function of temperature. Adapted with permission under a Creative Common CC BY-NC 4.0 license from refs (382). Copyright 2019 AAAS.
Figure 35
Figure 35
(a) Large-scale atomic-resolution STM topography (20 nm) of the 2H-MoTe2 surface. The image reveals an approximately uniform density of two types of defects over the entire surface. (b) DFT+U-optimized geometry for Mosub defect. (c) DFT+U-optimized geometry of the Mo vacancy Movac. (d) Magnetization density (0.001 electrons/bohr3) on the top surface of bulk 2H-MoTe2 in AF configuration. Spin-up and spin-down states are shown in faint blue and orange isosurfaces, respectively. Note that spins also couple antiferromagnetically at the local level between the Mo impurity and the nearest Mo atoms. All panels are adapted with permission under a Creative Common CC BY-NC 4.0 license from ref (382). Copyright 2019 AAAS.
Figure 36
Figure 36
Calculated magnetization of the antisite defect. (a) Occupation number versus rigid potential shifts α for antisite defects for the bare, noninteracting potential χ0 and the interacting potential χ. From the angular coefficients of both curves we can extract the optimum ULR for our system, ULR = formula image – χ–1. (b) Variation of the local magnetization at the defect antisite versus U. At U = 0, no magnetic moments are observed as the defect shows a symmetric configuration at the Mo–Mo bonds. At U > 0.5 eV, this symmetry is broken and the defect develops an appreciable magnetic moment that increases with U as a result of the increased localization of the bands. All panels are adapted with permission under a Creative Common CC BY-NC license from ref (382). Copyright 2019 AAAS.
Figure 37
Figure 37
Temperature dependence of the magnetic volume fraction for CrI3. Adapted with permission under a Creative Commons CC BY license from ref (383). Copyright 2021 Springer Nature.
Figure 38
Figure 38
(a) Temperature dependence of the magnetic volume fraction for VI3, determined from weak transverse field μSR experiments. (b) Zero-field μSR spectra, recorded at T = 5 and 60 K. (c) The temperature dependence of the internal fields Hint for VI3. Original figure, no permissions needed.
Figure 39
Figure 39
Voltage control of the magnetic properties of CrI3, CGT and FGT. (a) Top: Normalized magnetization measured by MCD as a function of the applied electric field (trace and retrace) at 4 K and fixed magnetic field (+0.44 T for top panel and −0.44 T for bottom panel), showing the electrical switching of the magnetic order in bilayer CrI3. The insets represent the corresponding magnetic states. Adapted with permission from ref (8). Copyright 2018 Springer Nature. (b) Uniaxial magnetic anisotropy field formula image of multilayer CGT as a function of temperature at different gate voltages and in the pristine case. Inset: The dependence of TC on gate voltage. Adapted with permission from ref (86). Copyright 2020 Springer Nature. (c) TC of a trilayer FGT as a function of gate voltage. (d) HC of a trilayer FGT as a function of gate voltage at 10 K. Panels (c) and (d) are adapted with permission from ref (12). Copyright 2018 Springer Nature.
Figure 40
Figure 40
Current-induced magnetization switching of a FGT 2D magnet. (a) Schematic illustration of the current-induced switching of a FGT nanoflake by the spin current generated by SHE in the Pt layer deposited on FGT and injected into FGT to produce a SOT.Jx is the current in Pt generating a downward spin current by SHE, H0 is the in-plane field tilting the magnetization M from its out-of-plane orientation at zero field, HDL is the damping-like (DL) component of the effective field expressing the action of the spin transfer torque. (b) Current-induced switching detected by the change of sign of the transverse voltage VH in panel (a) (RHVH/Jx). Panels (a) and (b) are adapted with permission from ref (442). Copyright 2019 American Chemical Society. (c) Schematic view of the CrI3/TaSe2 vdW heterostructure consisting of an insulating AF bilayer of CrI3 and a nonmagnetic metallic monolayer TMD TaSe2. The SOT on the magnetization of CrI3 is due to the charge-to-spin Edelstein conversion of the current flowing along the CrI3/TaSe2 interface. The resulting switching of m1 is indicated by arrows. Adapted with permission from ref (444). Copyright 2020 American Chemical Society.
Figure 41
Figure 41
Structure and magnetic properties of Bi2Te3/FGT heterostructures. (a) Structure of Bi2Te3/FGT. (b) Anomalous Hall resistance Rxy as a function of temperature for FGT and Bi2Te3/FGT heterostructures. (c) Magnetic phase diagram of pure FGT and Bi2Te3/FGT heterostructures versus FGT layer thickness and temperature (FM for ferromagnetic, PM for paramagnetic). All panels are adapted with permission from ref (90). Copyright 2020 American Chemical Society.
Figure 42
Figure 42
Spin Seebeck effect and magnon transport with 2D magnets. (a) Crystal structure of CGT and CST. (b) Schematic of the longitudinal SSE measurements in CST/Pt or CGT/Pt bilayers. H denotes the external magnetic field and ΔT (∇T) the temperature difference (gradient). (c, d) Normalized SSE voltage S = (V/∇T) (Lz/Ly) as a function of H in the (c) CST/Pt and (d) CGT/Pt bilayers at selected temperatures. Panels (a–d) are adapted with permission from ref (467). Copyright 2019 American Physical Society. (e) Schematic of the magnon generation, transport, and detection in MnPS3. (f) Optical image of the device with the MnPS3 flake and Pt electrodes, including the measurement configuration of the nonlocal SSE. (g) Normalized nonlocal signal formula image as a function of distance (d) for selected temperatures in a 16 nm-thick MnPS3 flake. The solid lines represent the best-fitting results based on a diffusion equation. (h) Magnon diffusion length as a function of MnPS3 thickness (t) for selected temperatures. Panels (e–h) are adapted with permission under a Creative Commons CC BY 4.0 license from ref (468). Copyright 2019 American Physical Society.
Figure 43
Figure 43
Skyrmions in 2D magnets. (a) Spin structure of a Néel skyrmion. (b) Pt/Co interface in which the absence of inversion symmetry generates DMI, HDMI = −(S1 × S2D12. (c) Side view of the crystal structure of the Janus TMD MnSTe in which the absence of inversion symmetry generates DMI. (d) DMI strength calculated for the Janus TMD MnSeTe, MnSTe, and MnSSe. (e) Skyrmions in MnSeTe (T = 10 K in applied field of 0.3 T) from DFT calculation of DMI and Monte Carlo simulations. Panels (c–e) are adapted with permission from ref (480). Copyright 2020 American Physical Society. (f) Side view of the crystal structure of WTe2 on FGT. (g) LTEM image of Néel skyrmion lattice at 180 K under 510 Oe in the sample 2L WTe2/40L FGT (L = layer) at tilt angle 30° and under focus. Scale bar: 500 nm. Panels (f) and (g) are adapted with permission under a Creative Commons CC BY license from ref (27). Copyright 2020 Springer Nature. (h) TEM image of Néel skyrmion lattice in an oxidized FGT flake (about 50 μm thick) at 160 K, tilt angle 20° and over focus. Adapted with permission from ref (26). Copyright 2021 American Physical Society. (i) Magnetization maps derived from analysis of LTEM image for Bloch bubbles in a CGT flake at 17 K in a field of 11.7 mT. Adapted with permission from ref (481). Copyright 2019 American Chemical Society.
Figure 44
Figure 44
Toward MRAM with 2D magnets. (a) Schematic of STT-MRAM with 3D materials: information coded by the relative orientations of the magnetization of the two magnetic layers (green) separated by an insulating MgO layer (red), writing by current-induced STT and reading by TMR. Adapted with permission from ref (491). Copyright 2017 American Physical Society. (b) Schematic of the TMR device based on a FGT/hBN/FGT vertical stack used for the results shown in panels (c–e). (c) TMR measurement in a FGT/hBN/FGT vertical stack at 4.2 K. The swapped magnetic field is out of plane. (d, e) Magnified regions of the TMR measurement around the field range of antiparallel configuration (upper panels) and variation of the AHE resistance of the top (blue) and bottom (green) FGT electrodes in the same field range. Blue and green arrows indicate the successive orientations of the magnetizations. Panels (b–e) are adapted with permission from ref (109). Copyright 2018 American Chemical Society. (f) Schematic of the device combining a gate-controlled spin-flip transition in bilayer CrI3 and spin filtering in the tunnel junction. Arrows indicate the magnetic orientation of each layer. (g) Tunnel conductance of the device illustrated in panel f as a function of gate voltage (sweeping back and forth) under a constant magnetic field (0.76 T). The measured tunnel conductance changes when the magnetic order of bilayer CrI3 is switched by the gate. Panels (f) and (g) are adapted with permission from ref (492). Copyright 2019 Springer Nature.
Figure 45
Figure 45
2D magnet-based SOT-MRAMs. (a) Top: Schematic of a SOT-MRAM based on 3D materials in which an electrical current in the heavy metal (Ta) of the bottom electrode generates by SHE the vertical spin current injected in the bottom FeCoB layer. This injection of spin current switches the magnetization of FeCoB by SOT (writing). The state of the memory is detected by the TMR of the FeCoB/MgO/FeCoB MTJ (reading). Bottom: Detection by TMR of the SOT-induced switching of the magnetization of the bottom FeCoB layer in the device of the schematic. Adapted with permission from ref (507). Copyright 2014 AIP Publishing. (b) Top: Schematic of a bilayer for SOT-MRAM in which the orientation of the out-of-plane magnetization of a FGT layer codes the information and is switched by the SOT generated by the SHE of the Pt layer. As shown in the bottom part of the figure, the switching is detected by the AHE resistance Rxy derived from the voltage between transverse contacts. Adapted with permission under a Creative Commons CC BY-NC 4.0 license from ref (443). Copyright 2019 AAAS. (c) Image of a heterostructure for SOT-MRAM in which the magnetic state of a CGT layer can be switched by the SHE of a Ta layer. (d) Comparison of the current densities and in-plane fields required for SOT switching in devices based on 3D magnetic layers (CoFeB, MnGa, thulium iron garnet (TmIG)) and 2D magnets (FGT, CGT). The best results so far are for Ta/CGT. Panels (c) and (d) are adapted with permission from ref (446). Copyright 2020 John Wiley and Sons.
Figure 46
Figure 46
Devices based on current-induced motion of skyrmions in 2D magnets. (a) Skyrmion motion induced by current pulses in a FGT track. Each STXM image is acquired after injecting five unipolar current pulses of 50 ns. Two individual Néel skyrmions are outlined in colored circled for clarity. The diameter of the skyrmions is about 200 nm, their velocity is around 1 m/s for a current density of 1.4 × 1011 A/m2 and the width of the track is 50 μm. Adapted with permission from ref (26). Copyright 2021 American Physical Society. (b) Schematic of racetrack memory storing data by aligning skyrmions like beads on an abacus and displacing them by current-induced SOT from write head to read head. Adapted with permission from ref (476). Copyright 2018 AIP Publishing. (c) Proposal of skyrmion-based racetrack memory based on the SOT-induced motion of antiferromagnetically-coupled skyrmions in two layers coupled by AF interactions. The left inset is a schematic of an antiferromagnetic (AFM or AF)-coupled nanotrack and the right inset represents AF-coupled skyrmions in Co/Ru/Co trilayers. As the AF-coupled skyrmions have the same chirality but opposite polarities, their motion has the advantage of being along the current direction (no Skyrmion Hall effect,,). Note that the racetrack memory of the schematic includes not only injector and detector but also an update/delete/insert. Adapted with permission from ref (513). Copyright 2018 IEEE. (d) AF-coupled CrI3 layers in a CrI3 bilayer. Adapted with permission from ref (9). Copyright 2018 Springer Nature.
Figure 47
Figure 47
(a) Magnetic structure of Co3Sn2S2, showing a FM ground state with spins on Co atoms aligned along the c-axis. (b) Kagome lattice structure of the Co3Sn layer. (c) Topographic image of the CoSn surface. (d) A zoom-in image of the CoSn surface (left) that shows similar morphology with the FeSn surface (right) in Fe3Sn2. The inset illustrates the possible atomic assignment of the kagome lattice. Panels (a–d) are adapted with permission under a Creative Commons CC BY 4.0 license from ref (514). Copyright 2020 Springer Nature. (e) The temperature dependence of the intrinsic anomalous hall conductivity in Co3Sn2S2. (f) Left: linear band crossings form a nodal ring in the mirror plane. Right: Spin–orbit coupling breaks the nodal ring band structure into opened gaps and Weyl nodes. The Weyl nodes are located just 60 meV above the Fermi level, whereas the gapped nodal lines are distributed around the Fermi level. Panels (e) and (f) are adapted with permission from ref (515). Copyright 2018 Springer Nature.
Figure 48
Figure 48
(a) Upper panel: Illustration of the magnetization-polarized Zeeman effect for Co3Sn2S2. Lower panel: illustration of the large negative orbital magnetism of the flat band in the kagome lattice. (b) Upper panel: Orbital magnetism for the flat band calculated from first principles. The magnetic moment (red arrows) is plotted along the flat band. The red bar marks the units of the magnetic moment value. Lower panel: Orbital magnetism from the magnetic kagome lattice model. The magnetic moment (red arrows, arbitrary unit) is plotted along the flat band. All panels adapted with permission from ref (516). Copyright 2019 Springer Nature.
Figure 49
Figure 49
(a) The temperature dependence of the relative volume fractions of the two magnetically ordered regions. Arrows mark the critical temperatures TC1 and TC2 for FM and AF (or AFM) components, respectively as well as the transition temperature formula image, below which only FM component is observed. (b) Spin structures of Co3Sn2S2, i.e., the FM and the in-plane AF (or AFM) structures. (c) The correlation plot of anomalous hall conductivity versus FM fraction. (d) Calculated AHC for out-of-plane FM and in-plane AF structures. The inset shows the calculated Berry curvature distribution in the BZ at the FM phase. All panels are adapted with permission under a Creative Commons CC BY 4.0 license from ref (514). Copyright 2020 Springer Nature.
Figure 50
Figure 50
(a) Calculated fractions of the ferromagnetism F and antiferromagnetism (1 – F) in Co3Sn2S2 and (b) averaged Chern number as a function of the in-plane AF correlation Jxy, taking into account fluctuations in the Hund’s coupling. All panels are adapted with permission under a Creative Commons CC BY 4.0 license from ref (535). Copyright 2020 American Physical Society.
Figure 51
Figure 51
Mechanical exfoliation methods. (A) Schematic of the Al2O3 film-assisted mechanical exfoliation. The strong adhesion between the crystal and the Al2O3 film makes it possible to exfoliate layered crystals that are otherwise difficult to cleave from SiO2 surfaces using conventional methods. Adapted with permission from ref (12). Copyright 2018 Springer Nature. (B) Schematic of the Au-assisted exfoliation process. First, a thin layer of Au is deposited onto a substrate, then a freshly cleaved bulk crystal is placed on the Au layer. The Au is then removed with a KI/I2 aqueous solution etchant. Adapted with permission under a Creative Commons CC BY license from ref (566). Copyright 2020 Springer Nature.
Figure 52
Figure 52
Identification of exfoliated layer numbers. (A) Optical contrast map of a representative CrI3 flake (left) and the corresponding optical contrast per number of layers (right). Adapted with permission from ref (5). Copyright 2017 Springer Nature. (B) Left: Optical image of few-layer flakes of MnBi2Te4 exfoliated onto Al2O3. The corresponding layer number is labeled on selected flakes. The scale bar is 20 μm. Right: MnBi2Te4 transmittance versus layer number. Adapted with permission from ref (111). Copyright 2020 AAAS. (C) Optical image of few-layer Fe3GeTe2 flakes exfoliated onto Al2O3. (D) AFM image of the area in (C) marked by a solid black square. The scale bar is 2 μm. (E) Height profile plotted versus length along the white line in (D). Panels (C–E) are adapted with permission from ref (12). Copyright 2018 Springer Nature.
Figure 53
Figure 53
Scheme for the one-step synthesis and vapor transport of CrX3 (X = Br, I) micro- and nanosheets directly on YSZ substrates shown by the example of CrBr3. Prior to the CVT process for deposition of the respective nanolayers the introduction of chromium powder and bromine (Br2 in small sealed capillaries) lead to the formation of CrBr3 (solid) (Cr(s) + 1,5 Br2(l) → CrBr3(s)) and gaseous CrBrn(g) (n = 2, 3, 4) at T2. By application of a temperature gradient (T2T1) chemical vapor transport is achieved for deposition of CrBr3 micro- and nanosheets on YSZ substrates directly. The reaction course is similar for the formation of CrI3, while in contrast to this scheme CrCl3 is utilized as presynthesized compound, that is not introduced by mixture of the elements. Reproduced with permission from ref (600). Copyright 2019 John Wiley and Sons.
Figure 54
Figure 54
Crystal growth by vapor transports of CrCl3 on YSZ substrates. (a) Optical microscopy of CrCl3 micro- and nanocrystals at YSZ substrate. (b) Optical microscope image of YSZ substrate surface with CrCl3 sheets with respective thicknesses in dotted boxes. (c) Distribution of thicknesses of CrCl3 structures on a YSZ substrate after CVT (yellow), after one time of exfoliation (green) and after three times of exfoliation (purple). (d) AFM measurement the height profile of a CrCl3 nanosheet. (e) Corresponding AFM image of measurement of (d) the white arrow is indicating the measurement. (f) AFM measurement of the height profile a CrCl3 ultrathin sheet (red line) and monolayer (purple line) after three repeats of exfoliation. (g) Corresponding AFM image of measurement of (f) the white line is indicating the monolayer AFM measurement. All panels are reproduced with permission from ref (600). Copyright 2019 John Wiley and Sons.
Figure 55
Figure 55
(A, B) RHEED patterns with indicated diffraction orders of (A) the bare HOPG substrate and (B) the MBE-grown CrBr3 film. (C, D) STM images of (C) the CrBr3 monolayer with (D) bilayer islands. The scan parameters were as follows: Vb = 1.1 V, I = 100 pA, T = 5 K for (C) and Vb = 1.5 V, I = 100 pA, T = 5 K for (D). (E) Atomically resolved image of a monolayer CrBr3 with an overlaid atomic structure. The scan parameters were as follows: Vb = 1.5 V, I = 500 pA, T = 5 K. The lattice constants were determined to be 6.3 Å for the primitive vectors a and b, consistent with the bulk values. (F) Illustrations of the top and side views of the monolayer CrBr3 atomic structure. The Cr atoms form a honeycomb lattice sandwiched by Br atoms. Within the Cr honeycomb lattice, the top and bottom surfaces of Br atoms form single triangles but with opposite orientation, indicated by solid and dotted green lines, respectively. (G) AFM image of monolayer CrBr3 with partial coverage. A line-cut profile across the monolayer and bare substrate is shown with a monolayer height of 6.5 Å. All panels are adapted with permission from ref (20). Copyright 2019 AAAS.
Figure 56
Figure 56
(A, B) The structure and ground states of iron-deficient Fe3–xGeTe2. (A) Side view of stoichiometric Fe3GeTe2. FeI (red) and FeII (silver) are two inequivalent Fe sites with +3 and +2 formal charges, respectively. The FeI–FeI interactions are mostly responsible for interlayer AF ordering while FeI-FeII and FeII–FeII couplings are FM. With Fe defects or doping, FeI–FeII and FeII–FeII become dominant and push interlayer ordering into FM. (B) The calculated energy differences (ΔE) between interlayer AF and FM phases as a function of hole concentration. For ΔE > 0, FM is favored (between 0.2 and 0.6 holes per formula unit). Panels (A) and (B) are adapted with permission from ref (615). Copyright 2020 American Chemical Society. (C) The layered crystal structure of MBT. The red arrow indicates the Mn sites in which antisites have been identified. (D) STM of the surface of a cleaved MBT crystal; white spots show the presence of multiple antisite point defects. Panels (C) and (D) are adapted with permission from ref (616). Copyright 2020 American Chemical Society. (E) Crystal structure representation of the (MnBi2Te4)m(Bi2Te3)n series, ranging from AF to FM with various compositions derived from codepositional MBE. Here, we see that with the addition of Bi2Te3 layers, the MnBi2Te4 layers cannot be coupled together and the material becomes less AF. This is a prime example of an off-stoichiometry defect. Panel (E) adapted with permission from ref (617). Copyright 2020 AIP Publishing.
Figure 57
Figure 57
Common experimental techniques for measuring the Young’s modulus and strength of 2D materials. (a) Nanoindentation based on AFM. Adapted with permission under a Creative Commons CC BY license from ref (634). Copyright 2019 John Wiley and sons. (b) Bulge test involving interferometry and Raman spectroscopy. Adapted with permission from ref (635). Copyright 2017 American Physical Society. (c) Tensile testing push-to-pull micromechanical device controlled by an external pico-indenter in SEM, and the yellow arrow shows the loading direction. (d) Diagram showing the enlarged pink rectangle area in (c) with a suspended graphene sample after cutting by focused ion beam., Copyright 2009 Royal Society of Chemistry, copyright 2019 American Physical Society, copyright 2019 John Wiley and Sons, copyright 2017 American Physical Society, copyright 2020 Springer Nature, copyright 2014 Springer Nature. Panels (c) and (d) are adapted with permission under a Creative Commons CC BY 4.0 license from ref (636). Copyright 2020 Springer Nature.
Figure 58
Figure 58
Mechanical properties of CrCl3 and CrI3. (a) Optical microscopy images of 2L and few layers CrCl3 and Crl3 suspended over microwells (600 nm in diameter) on a SiO2/Si substrate. (b) Load–displacement curves and the corresponding fittings of 2L and 7L CrCl3 and CrI3. (c) Volumetric Young’s modulus and breaking strength of 2–10L CrCl3 and CrI3, along with dashed lines that show the linear fits. (d) Demonstration of the good plasticity of bulk CrCl3 and CrI3 crystals via folding them into rings. (e) Deformability factor versus Young modulus, where I, II, and III correspond to plastic-flexible, potentially deformable and brittle-rigid regions, respectively. The experiential results of CrCl3 and CrI3 are shown as red-filled circles, and the other layered vdW materials are shown as green filled circles. (f) Deformability factor versus bandgap for the same materials as in (e), and the materials that may show exceptional plastic behavior are shown in the dashed line encircled green area. All panels are adapted with permission from ref (659). Copyright 2021 American Chemical Society.
Figure 59
Figure 59
Modulus-strength graph. The Young’s modulus and fracture strength of 2D magnets are compared with those of conventional bulk materials and other 2D materials. Adapted with permission from ref (659). Copyright 2021 American Chemical Society.
Figure 60
Figure 60
Mechanical properties of ternary compounds. (a) Changes in the magnetic configurations of various MPX3 (M = Mn, Fe, Ni; X = S, Se) compounds at zero carrier density under in-plane biaxial compressive and expansive strains. Adapted with permission from ref (684). Copyright 2016 American Physical Society. (b) Magnetostriction in 2L CrI3 and (c) 6L CrI3 resonators under an out-of-plane magnetic field (μ0H) and in (d) 6L CrI3 resonator under an in-plane magnetic field (μ0H). The resonance frequency (b–d) and MCD (b, c) of the membranes as a function of the magnetic field, where the red (blue) lines correspond to the measurement for the positive (negative) sweeping direction of the field. Panels (b–d) are adapted with permission from ref (685). Copyright 2020 Springer Nature.
Figure 61
Figure 61
Magnetic order and spin excitations in vdW magnetic materials with tunable fundamental spin Hamiltonian’s and structural parameters probed by neutron scattering methods. Adapted with permission from ref (84). Copyright 2018 Springer Nature.
Figure 62
Figure 62
Magnetic structures of (a) FePS3. Adapted with permission from ref (724). Copyright 2016 American Physical Society. (b) MnPS3. adapted with permission from ref (318). Copyright 2006 American Physical Society. (c) NiPS3. adapted with permission from ref (317). Copyright 2015 American Physical Society. (d) CoPS3. adapted with permission from ref (729). Copyright 2017 IOP Publishing. All of these materials have honeycomb lattice structure and are antiferromagnetically ordered.
Figure 63
Figure 63
(a) A spin gap at Dirac point in CrI3 suggests that spin excitations in this system can have chiral and topological edge mode. Adapted with permission under a Creative Commons CC BY 4.0 license from ref (744). Copyright 2018 American Physical Society. (b) The FM phase transition in CrI3 is weakly first-order and controlled by spin gap. Adapted with permission from ref (745). Copyright 2020 American Physical Society.
Figure 64
Figure 64
(a) The phase diagram of α-RuCl3 adapted with permission from ref (752). Copyright 2018 Springer Nature. At zero field, the system forms a zigzag magnetic structure as shown in the left inset. For in-plane magnetic field between 7 and 9 T, the system is believed to be in Kitaev QSL state. For fields above 9 T, the system becomes non topological from thermal transport measurements. (b) Wave vector/energy dependence of spin excitation of α-RuCl3 at 4 and 8 T, adapted with permission under a Creative Commons CC BY 4.0 from ref (753). Copyright 2018 Springer Nature. One can clearly see spin waves stemming from zigzag ordered wave vector (0.5, 0, 0) at 4 T. The scattering centered around Γ points is believed to arise from fractionalized excitations of a Kitaev QSL, which is enhanced upon suppression of spin waves with a 8 T in-plane magnetic field.
Figure 65
Figure 65
(a) The 2D kagome lattice structure. (b) Calculated electronic dispersion and flat band. (c, d, e) Crystal structures of T3X, T3X2, and TX, respectively, where T = Fe, Mn, Co, and X = Ge, Sn. All panels are adapted with permission from ref (766). Copyright 2019 Springer Nature.
Figure 66
Figure 66
(a) Red arrows indicate in-plane spin directions at zero field. Black arrows indicate the moment direction in a c-axis aligned field. (b) Spin-wave branches for one acoustic and two optical bands. The middle and top green arrows indicate Dirac points and band top, respectively. (c) Nearest-neighbor DMI direction. (d) DMI induced spin gap at Dirac points. (e) Neutron scattering measured spin waves energy/moment map at (a) 0 T, (b) 2 T, (c) 7 T; (d) Calculated neutron spectra for 7 T. All panels are adapted with permission from ref (767). Copyright 2015 American Physical Society.
Figure 67
Figure 67
(a) A-type AF structure of FeSn in 2D kagome lattice structure, adapted with permission from ref (779). Copyright 2019 American Physical Society. (b) The antichiral noncollinear structure of Mn3Ge, adapted with permission from ref (673). Copyright 2020 American Physical Society. (c, d) Complicated noncollinear magnetic structures of barlowite with proximate 2D kagome lattice structure, adapted with permission under a Creative Commons CC BY 4.0 license from ref (780). Copyright 2020 Springer Nature.
Figure 68
Figure 68
(a) Top: Schematic illustrating IETS mechanism. Bottom: Color plot of IETS spectra (|(d2I)/(dV2)| versus V) taken on bilayer CrI3 as a function of out-of-plane magnetic field (easy axis) shows two dispersing magnon modes. Bottom panel adapted with permission from ref (120). Copyright 2019 National Academy of Sciences of the United States of America. (b) Top: Spin-wave calculations of magnon dispersion in 2D CrI3 from anisotropic Heisenberg model with nearest-neighbor interactions. Adapted with permission from 217. Copyright 2018 Springer Nature. Magnon energies are shown at high-symmetry Γ and M points as a function of nearest-neighbor exchange energy J and anisotropy α. Zoom-in of the acoustic branch at the Γ point shows an energy gap for α > 1. (c) Extracted J and α values for all three 2D chromium trihalides from IETS measurements.
Figure 69
Figure 69
(a) Top: Experimental Raman spectra (left) and theoretical calculations (right) of magnon modes in bulk CrI3 as a function of out-of-plane magnetic field (easy axis). Above Bc ∼ 2 T, a single mode is observed as expected for a fully spin-polarized state for all layers. Below Bc, two additional modes are seen that disperse oppositely with field, corresponding to the layer-AF order on surface layers. Bottom: Schematic of spins in bulk CrI3 above and below Bc. Adapted with permission under a Creative Commons CC BY 4.0 license from ref (218). Copyright 2020 American Physical Society. (b) Field-dependent Raman spectra of bilayer CrI3. Only layer-AF magnons modes are obtained. Adapted with permission from ref (794). Copyright 2020 Springer Nature.
Figure 70
Figure 70
Process flow for vdW heterostructure fabrication. Schematic of the dry-polymer-transfer process used to fully encapsulate a 2D flake with hBN (here, it is graphene). This process relies only on the vdW interactions between hBN and other 2D flakes and can be used to pick up and transfer a number of different 2D materials. Reproduced with permission from ref (577). Copyright 2013 AAAS.
Figure 71
Figure 71
Contact methods for exfoliated vdW magnets. (A) Schematic of a Cr2Ge2Te6 flake top contacted with metal electrodes under an ion-liquid gate. Adapted with permission from ref (86). Copyright 2020 Springer Nature. (B, C) Optical image of a 5.8 nm-thick Fe3GeTe2 (B) and a 28 nm-thick Fe5GeTe2 (C) flake contacted from the bottom with prepatterned metal electrodes. In (B) the red dashed line is the Fe3GeTe2 and the yellow dashed line is the hBN. The red scale bar is 10 μm. In (C) the scale bar is 20 μm. Panel (B) was adapted with permission from ref (123). Copyright 2018 Springer Nature. Panel (C) was adapted with permission from ref (87). Copyright 2019 American Chemical Society. (D–G) Schematics (D, F) and corresponding false-colored optical images (E, G) of CrI3 flakes contacted by graphene electrodes for lateral (D, E) and tunneling (F, G) transport measurements. In (E) and (G), the scale bars are both 5 μm. Panels (D–G) are reproduced with permission from ref (15). Copyright 2018 Springer Nature.
Figure 72
Figure 72
Heterostructures fabricated from 2D vdW magnets. (A) Schematic of a Cr2Ge2Te6 flake fully encapsulated with hBN and contacted by graphene electrodes. Reproduced with permission from ref (10). Copyright 2018 Springer Nature. (B) Schematic (left) and corresponding optical image (right) of a MTJ fabricated from Fe3GeTe2 electrodes separated by an hBN barrier. In the right panel, the dotted lines outline the edges of the two Fe3GeTe2 flakes. The scale bar is 5 μm. Reproduced with permission from ref (109). Copyright 2018 American Chemical Society. (C) Optical image (left) and corresponding cartoon (right) of a spin-filter MTJ utilizing CrI3 as the tunnel barrier between graphene electrodes. Reproduced with permission from ref (14). Copyright 2018 AAAS. (D) Schematic (top) and a false-colored optical image (bottom) of a spin-field-effect transistor fabricated from 4-layer CrI3. Graphene acts as both transistor electrodes and local electrostatic gates. The scale bar is 5 μm. Reproduced with permission from ref (504). Copyright 2019 American Chemical Society. (E, F) Schematic (E) and false-colored optical image (F) of a heterostructure proximitizing CrI3 with WSe2. The scale bar in (F) is 5 μm. Panels (E, F) are reproduced with permission under Creative Commons CC BY-NC 4.0 license from ref (206). Copyright 2017 AAAS.
Figure 73
Figure 73
2D magnetism controlled by twistronics and stacking order. (A) Variation of stacking order in small twist angle θ monolayer–bilayer graphene (tMBG). (B) Transverse resistance Ryx map measured versus total carrier density n and the magnetic field B at T = 6.4 K in a tMBG device with θ = 1.25° near ν = 3 orbital magnetic state presents a magnetization reversal driven by the change in n or B due to the non-negligible contribution from the topological edge states in large moiré unit cell area. ν = nA is the number of carriers n per moiré unit cell A. (C) In B-field, ν = 3 state switches between K and K′ valley polarization as doping level changes across the gap. (D) Nonvolatile switching between K and K′ valley-polarized magnetic states independently controlled by either n or B. Panels (A–D) reproduced with permission from ref (828). Copyright 2020 Springer Nature. (E) Temperature-dependent hysteresis (highlighted by colored areas) observed in magnetic field in rhombohedral graphite. (F) Phase diagram of the critical behavior in (E) characteristic to strongly correlated electronic systems. Panels (E, F) are reproduced with permission from ref (832). Copyright 2020 Springer Nature. (G) Scanning tunneling spectroscopy map of a small θ double bilayer graphene showing Bernal (black) and rhombohedral (white) stacking domains. Reproduced with permission under a Creatice Commons CC BY-NC 4.0 license from ref (833). Copyright 2021 National Academy of Sciences. (H–J) Rhombohedral (R) and monoclinic (M) stacking configuration and magnetic domains in a small θ twisted bilayer CrI3. (H) Three types of stacking domain walls in this system. Arrows represent the stacking vectors. (I) Sketch of the magnon network at θ = 0.1°. (J) Stacking and magnetic domain patterns of the gray rectangle area in (I). Red (blue) arrows represent stacking vectors for R (M) stacking. Red (cyan) lines represent the RR (MM) stacking domain walls. Panels (H–J) are reproduced with permission from ref (834). Copyright 2020 American Physical Society.
Figure 74
Figure 74
Tuning magnetic vdW heterostructures. (A) Schematic of a high-pressure setup for a MTJ. Yellow line represents electrical leads. (B) Tunnel current It versus magnetic field H at pressures from 0 to 2.7 GPa in a bilayer CrI3 MTJ. Insets show spin alignments and optical image of the MTJ. Panels (A, (B) are reproduced with permission from ref (22). Copyright 2019 Springer Nature. (C) Proposed vdW heterostructure where exchange (EX) and spin–orbit (SO) coupling can be swapped by an electric field. Cr2Ge2Te6 magnetization is denoted by red arrows. Adapted with permission from ref (840). Copyright 2020 American Physical Society. (D, E) Dependence of RMCD on magnetic field in heterostructures of WSe2 and trilayer (D)/bilayer (E) CrI3 (shown in the insets). Panels (D, E) are reproduced with permission from ref (203). Copyright 2020 Springer Nature. (F–H) Photoemission and spin-dependent charge transfer in hBN encapsulated CrI3/WSe2 heterostructure. (F) A schematic of the heterostructure deposited on nanopillar array. (G) Spin-dependent charge transfer from spin-polarized states in WSe2 to CrI3 results in the highly p-doped WSe2, where an exciton can be turned into a localized charged trion (0D X+) via a hole capture process (H). Arrows in red and green (blue) denote the spin direction in WSe2 (CrI3). Panels (F–H) are reproduced with permission under a Creative Commons CC BY license from ref (841). Copyright 2020 Springer Nature.
Figure 75
Figure 75
Comparison of different Monte Carlo models calculating the Curie temperature of CrI3. (a) Classical 2D Heisenberg. (b) Classical Ising model. (c) Quantum-like simulation with temperature rescaling. All panels are adapted with permission from ref (865). Copyright 2015 American Physical Society.
Figure 76
Figure 76
Slater-Pauling or volcano plot for 2D magnets. (a) High-throughput screening undertook over several crystal structures and elements of the periodic table including formulas MX′2, MX, MX′3, MPX3, MX2, and CrFTe3 with M = Sc–Zn, La, Y; X′=Cl, Br, I; X=O, S, Se, Te; F=Si, Ge. The simulations included mainly transition metals with 3d electrons, but some with 4d and 5d were included for comparison. (b) Variation of the local magnetic moment MB) at the metal atom as function of its valence Z(e). Bader charge analysis was used to extract Z for each metal atom at the compound. The solid lines show a fit to the data set on two different regimes according to the filling of the valence. The positive slope (weak magnets) can be fairly well fitted using M+ = 0.84Z – 1.15 (with a linear regression coefficient R2 = 0.96) and the negative (strong magnets) with M = −0.87Z + 9.27 (R2 = 0.90). An electron counting argument can be used to explain both regimes as discussed in the text. (c) Spin resolved density of states (DOS) for monolayer MPTe3 (M = V, Cr, Mn, Fe, Co, Ni) as function of the energy ε displaying the spin up density nup (faint gray) and spin down ndown (faint brown) at opposite sides. The energy is shifted to the Fermi energy εF at zero. (d) Variation of the model predicted magnetization versus DFT + U calculated magnetization for the compounds showed in (a). Calculations were performed using the VASP code using a 21 × 21 × 1 k-sampling grid, the Dudarev (GGA+U) scheme with Hubbard U values following those in ref (341). The energy cutoff is set to 600 eV, the convergence criteria for energy to 10–7 eV and for the forces to 0.01 eV/Å. In order to avoid interactions between the layers, we applied periodic boundary conditions with a vacuum space of 25 Å. We used the projector augmented wave (PAW) methods with a plane wave basis. The Vosko–Wilk–Nusair modification scheme is applied for the spin-polarized calculations. All images in this figure are original, and no permissions are required.
Figure 77
Figure 77
Comparison of spin Hamiltonians used to model the magnetic properties of CrI3. The two most commonly used Hamiltonians in the literature have been the Ising and the Heisenberg (XXZ) models which also includes magnetic anisotropy. The latter was used to understand inelastic tunnelling spectra for MTJs. The former was initially assigned to CrI3 but its gross overestimation of the Curie temperature relative to experimental data made it unrealistic to account for the interactions in the system. Overall, depending on the property being measured, other alternatives have been considered: (i) for angle-dependent FM resonance measurements, the Kitaev model with quadrupole–quadrupole interactions and the Zeeman coupling was implemented; (ii) for inelastic neutron scattering to extract the magnon dispersion of bulk CrI3,, the XXZ model either including DMI or adding biquadratic exchange with DMI have been proposed; and (iii) for the magnetic domains and domain walls on CrI3, a Hamiltonian taking into account biquadratic exchange was utilized. Starting with image at top, and going clockwise, panels adapted with permission under a Creative Commons CC BY license from refs (796). Copyright 2020 Springer Nature. Adapted with permission from ref (14). Copyright 2018 AAAS. Adapted with permission from ref (745). Copyright 2020 American Physical Society. Reproduce with permission from ref (31). Copyright 2021 John Wiley and Sons. Adapted with permission from ref (746). Copyright 2018 AAAs.
Figure 78
Figure 78
Merons and antimerons on 2D magnet CrCl3. (a) Artistic view of the presence of topologically nontrivial spin quasiparticles merons and antimerons on monolayer CrCl3. The different colors follow the orientation of the spins throughout the layer. (b, c) Local views of antimerons and merons, respectively, with their local spin configurations at Cr atoms.

References

    1. Du K.-Z.; Wang X.-Z.; Liu Y.; Hu P.; Utama M. I. B.; Gan C. K.; Xiong Q.; Kloc C. Weak van der Waals Stacking, Wide-Range Band Gap, and Raman Study on Ultrathin Layers of Metal Phosphorus Trichalcogenides. ACS Nano 2016, 10, 1738–1743. 10.1021/acsnano.5b05927. - DOI - PubMed
    1. Kuo C.-T.; Neumann M.; Balamurugan K.; Park H. J.; Kang S.; Shiu H. W.; Kang J. H.; Hong B. H.; Han M.; Noh T. W.; Park J.-G. Exfoliation and Raman Spectroscopic Fingerprint of Few-Layer NiPS3 van der Waals Crystals. Sci. Rep. 2016, 6, 20904.10.1038/srep20904. - DOI - PMC - PubMed
    1. Lin M.-W.; Zhuang H. L.; Yan J.; Ward T. Z.; Puretzky A. A.; Rouleau C. M.; Gai Z.; Liang L.; Meunier V.; Sumpter B. G.; Ganesh P.; Kent P. R. C.; Geohegan D. B.; Mandrus D. G.; Xiao K. Ultrathin Nanosheets of CrSiTe3: A Semiconducting Two-Dimensional Ferromagnetic Material. J. Mater. Chem. C 2016, 4, 315–322. 10.1039/C5TC03463A. - DOI
    1. Lee S.; Choi K.-Y.; Lee S.; Park B. H.; Park J.-G. Tunneling Transport of Mono- and Few-Layers Magnetic van der Waals MnPS3. APL Mater. 2016, 4, 086108.10.1063/1.4961211. - DOI
    1. Huang B.; Clark G.; Navarro-Moratalla E.; Klein D. R.; Cheng R.; Seyler K. L.; Zhong D.; Schmidgall E.; McGuire M. A.; Cobden D. H.; Yao W.; Xiao D.; Jarillo-Herrero P.; Xu X. Layer-Dependent Ferromagnetism in a van der Waals Crystal Down to the Monolayer Limit. Nature 2017, 546, 270–273. 10.1038/nature22391. - DOI - PubMed

Publication types